DOI:
10.1039/C5RA05220C
(Paper)
RSC Adv., 2015,
5, 55396-55400
Towards free-standing MoS2 nanosheet electrocatalysts supported and enhanced by N-doped CNT–graphene foam for hydrogen evolution reaction†
Received
24th March 2015
, Accepted 10th June 2015
First published on 10th June 2015
Abstract
Large-scale free-standing porous carbon-based catalyst supports are critically needed for the hydrogen evolution reaction (HER) in view of their practical application. In this work, MoS2 nanosheets were uniformly deposited onto N-doped carbon nanotube–graphene (N-CNT–G) hybrids, forming a three-dimensional (3D) free-standing architecture. The designed 3D hybrid materials exposed the MoS2 nanosheet catalyst on the one-dimensional CNTs and there was facilitated ion transportation because of the porous graphene foam, and these 3D hybrid materials were used as electrocatalysts for HER directly without any transferring process. Moreover, N doping decreased the overpotential to 0.08 V, which increased the stability and promoted the catalytic activity, providing a facile route to design advanced high-performance HER catalysts towards a free-standing configuration.
1. Introduction
Hydrogen is a well-accepted clean alternative to the hydrocarbon fuels used today if it can be sustainably produced from water-splitting.1,2 Electrochemical hydrogen evolution is a possible approach to meet this standard and has attracted increasing attention in recent years.3,4 But one urgent need in the hydrogen evolution reaction (HER) is to find non-noble metal materials to replace the commonly used expensive and rare Pt catalyst.5,6 Recently, molybdenum-based compounds have received more attention, such as Mo2C,7,8 MoP,9,10 MoB11 and so on. These Mo-based catalysts have very good HER activity. Research has also revealed that molybdenum sulfide (MoS2) has outstanding electrocatalytic properties for the HER.12–22 The active sites in MoS2 are found to rely on its structure, these active sites come from the edge of the sheets in the crystalline platelets whereas they come from unsaturated sulfur atoms on the surface of amorphous MoS2. However, the edge sites of MoS2 are more catalytically active than the MoS2 basal planes.4,15,16 Both computational and experimental results confirmed that the HER activity stemmed from the sulfur edges of the MoS2 plates.16,18,22,25 In the water-splitting processes, the uncoordinated sulfur atoms denote active sites that bind hydrogen protons and release H2 molecules.17 To improve the HER activity, exposing more of the surface and more edges of the MoS2 using nanoparticles, nanosheets or porous structures allows the formation of a more active electrocatalyst.17–19 To further improve its distribution and conductivity, the MoS2 catalysts were used with carbon nanomaterials, such as carbon nanotubes (CNTs) and graphene as a support,20–22 just like other types of electrocatalysts. Hence some well-established strategies, like using a CNT forest to increase the area-specific activity and applying N-doping to facilitate active species immobilization, were also suitable to produce advanced MoS2-based electrocatalysts.23,24 The recent study by Kim et al. has revealed that MoSx/N-doped CNT (NCNT) forest hybrid catalysts exhibited the highest HER activity of any MoS2-based catalyst ever reported.25 Based on this progress and considering their future application, a further interest in this field is exploring MoS2/heteroatom-doped CNT hybrid catalysts and especially free-standing architectures. Our previously developed CNT–graphene hybrid foams have been proven to form free-standing structures with high conductivity and large surface area, and rich in pores,26,27 demonstrating the excellent characteristics of an ideal catalyst support. In this study, free-standing three-dimensional (3D) N-doped CNT–graphene (N-CNT–G) hybrid foams were fabricated and used as a support to grow MoS2 nanosheets. The MoS2/N-CNT–G electrocatalysts showed highly efficient electrocatalysis towards the HER and kept an unbroken 3D framework during the hydrogen evolution process due to its free-standing structure.
2. Experimental section
2.1. Synthesis of 3D non-doped and N-doped CNT–G hybrid foams
Non-doped CNT–G hybrid foams were synthesized as described in our previous work.26,28 For the preparation of the N-CNT–G hybrid foams, pyridine instead of ethanol and benzene was used as the precursor to grow the N-doped graphene and CNTs. Here the remaining metal catalyst and Ni foam templates encapsulated in the graphene framework were removed by 3 mol L−1 HCl solution at 90 °C for 10 h.
2.2. Synthesis of free-standing MoS2/N-CNT–G hybrids
A piece of the N-CNT–G hybrid foam (∼5 mg) was carefully put into a Teflon-lined autoclave containing a preset amount of ammonium thiomolybdate ((NH4)2MoS4) and 10 ml of N,N-dimethylformamide. The mass ratio of N-CNT–G to (NH4)2MoS4 was controlled as 5
:
1, 1
:
3, 1
:
6 and 1
:
10. The reaction was carried out at 200 °C for 10 h and then cooled to room temperature naturally. Finally, the black MoS2/N-CNT–G hybrid foam was brought out and dried in the air after being washed with ethanol and de-ionized water in turn. For comparison, MoS2 was also deposited onto the CNT–G foam (termed as MoS2/CNT–G) through a similar process.
2.3. Characterization
The products were characterized by scanning electron microscopy (SEM, Hitachi S-4800) and transmission electron microscopy (TEM, JEOL-JEM-2100F), X-ray diffraction (XRD, Philips X’pert Pro X-ray diffractometer with Cu Kα radiation of 1.5418 Å), X-ray photoelectron spectroscopy (XPS, PHI5000 VersaProbe) and Raman spectroscopy (Renishaw inVia Raman Microscope with an argon-ion laser at an excitation wavelength of 514 nm).
2.4. Electrochemical measurements
The electrochemical performance of the catalyst was evaluated by linear sweep voltammetry (LSV) and cyclic voltammetry (CV). All electrochemical measurements were conducted at room temperature in a three-electrode cell connected to a CHI 660C workstation (CH Instrument, Inc.). A slice of the free-standing MoS2/N-CNT–G hybrid foam was bonded to a platinum wire to perform as the working electrode. A saturated calomel electrode and a graphite rod served as the reference and counter electrodes, respectively. The LSV was measured at a scan rate of 5 mV s−1 in a 0.5 M H2SO4 solution.
3. Results and discussion
The N-doped graphene and CNTs were readily grown on Ni foams in turn by chemical vapor deposition. SEM images, shown in Fig. 1a and b, revealed a similar morphology to our previously reported non-doped graphene and CNT–graphene foams.26 The presence of graphene and CNTs was revealed by the TEM images shown in Fig. 1c and S2.† As observed, the obtained multi-walled CNTs had diameters ranging from 20 to 60 nm. The content of the N dopant was about 1.06 at% according to the XPS analysis. It was mainly composed of pyridinic and pyrrolic N entities, which were depicted by the fitting peaks centered at 398.1 and 400.5 eV in Fig. 1d.28,29 In order to obtain a freestanding N-CNT–G foam, the Ni foam templates encapsulated in the graphene framework were removed by HCl. As demonstrated in the SEM image of the N-CNT–G foam provided in Fig. S1,† the hybrid film still kept a free-standing feature in a regular foam structure, although some broken junctions were created when the Ni foam template was removed.
 |
| Fig. 1 (a) SEM image of the N-doped graphene/Ni foam; (b) SEM image of the N-CNT–G/Ni foam; (c) TEM image of the detached N-CNT and graphene sheets; (d) N 1s XPS narrow scan of the N-CNT–G/Ni foam. | |
Fig. 2a shows the typical SEM image of the MoS2/N-CNT–G hybrid. The MoS2/N-CNT–G hybrid has the same free-standing structure as the N-CNT–G foam at low magnification (Fig. S1a†). After MoS2 deposition (Fig. 2b), the surface of the N-CNTs was uniformly coated by a layer of stacked nanosheets. The MoS2/N-CNT composites inherited the one-dimensional morphology from the nanotubes while their diameter increased significantly. The morphology of the MoS2/N-CNT–G was studied by TEM (Fig. 2c and S2†). The heavy MoS2 coating made the N-CNTs indistinguishable under TEM observation. The MoS2 nanosheets grew vertically out from the N-CNTs, and their thickness was estimated to be about 50–70 nm by comparing the diameter change between the N-CNTs (Fig. 1c) and MoS2/N-CNT. It can be observed that each nanosheet was composed of 6–10 MoS2 layers with a thickness of ∼8 nm (Fig. 2d). The interplanar distance of the lattice fringes was approximately 0.63 nm, corresponding to the (002) plane of MoS2.30–32 When deposited onto non-doped CNT–G, the MoS2 also maintained a similar morphology (Fig. 2e and f) and structure to MoS2/N-CNT. The MoS2 layer on the CNTs, however, is thinner than that on the N-CNTs, suggesting that N doping is favourable for MoS2 precursor adsorption and deposition.
 |
| Fig. 2 SEM (a and b) and TEM (c and d) images of the MoS2/N-CNT–G hybrid. SEM (e) and TEM (f) images of the MoS2/CNT–G hybrid. The inset in (d) is a locally enlarged TEM image of a MoS2 nanosheet. | |
The XRD patterns of the MoS2/N-CNT–G and MoS2/CNT–G hybrids in Fig. 3a display two diffraction peaks at 31.9° and 56.2° that are attributed to the (100) and (110) planes of MoS2.33,34 The broad peaks indicated that the nanosized or even amorphous MoS2 was synthesized by a hydrothermal method, which was very consistent with the TEM result (Fig. 2d). The other two peaks at 26.3° and 41.4° corresponded to the (002) and (100) planes of the graphite layers in the CNT walls. The characteristic features of the MoS2 were also reflected in the Raman spectra (Fig. 3b), where the two bands at 384.2 and 408.7 cm−1 were assigned to the in-plane (E12g) and out-of-plane (A1g) vibrational modes of the few-layer MoS2.35,36 In addition, the two spectra both gave two prominent peaks at 1340 and 1578 cm−1, ascribed to D band and G band of the CNTs. The intensity ratio of the D band to the G band (ID/IG) in each sample was different. The N-CNT based sample showed a higher ID/IG because the N doping produced a lower graphitic crystalline structure.28
 |
| Fig. 3 XRD patterns (a) and Raman spectra (b) of the MoS2/CNT–G and MoS2/N-CNT–G hybrids. | |
The composition and chemical state of the MoS2 on the N-CNT–G foam was analyzed by XPS and is shown in Fig. 4. The Mo 3d XPS curve in Fig. 4a can be deconvoluted into two categories of Mo species. One is the dominant Mo(IV) that raised signal peaks of 3d5/2 and 3d3/2 at 228.8 and 231.9 eV, respectively. Another is Mo(VI), possibly from MoO3 or MoS3, as reflected by the fitted shoulder peaks in Fig. 4a. The fitted S 2s peaks (Fig. 4b) suggested that the S entities in our sample consisted of a majority of divalent sulfide ions (S2−) and a little bridging S22−, suggesting the existence of unsaturated sulfur atoms.15,21,37 The mole ratio of S to Mo was measured to be 2.6, a little higher than the stoichiometric ratio of MoS2, further supporting the existence some uncoordinated sulfur atoms. Hence the XPS analysis results given in Fig. 4a suggested that the obtained MoSx should have been a mixture of dominant MoS2 and a small amount of MoS3 or MoO3. Because the unsaturated sulfur atoms were inclined to gather at the edge sites of the as prepared MoSx nanosheet, the S-rich structure would have given a negatively charged surface of molysulfide to facilitate proton adsorption and reduction to H2 via electrostatic absorption of positively charged H+. The XPS spectra of Mo 3d and S 2p corresponding to the MoS2/CNT–G and MoS2/N-CNT–G (Fig. S4†) hybrids confirmed that the valence state of Mo and S within both compounds were +4 and −2, respectively. However, after N-doping, the binding energy of S 2p1/2 and 2p3/2 was slightly downshifted by 0.3 and 0.2 eV, respectively. As proposed by Kim et al.,25 the S elements were principally connected with the N-CNTs. Hence the downshift of the S signal should have been caused by the protonation of lone-pair electrons from the N-doping sites.
 |
| Fig. 4 Mo 3d (a) and S 2p (b) XPS spectra of the MoS2/N-CNT–G hybrid. | |
The proceeding results indicated that the free-standing MoS2/N-CNT–G hybrids with S-rich MoS2 nanosheets and conductive CNT–G frameworks were successfully prepared. The electrochemical HER tests were performed using a three-electrode setup in 0.5 M H2SO4 solution. Here only half of the free-standing MoS2/N-CNT–G foam was immersed into the electrolyte (Fig. 5a) in order to avoid the effect of the Pt-wire conductor because Pt has a high activity for the HER. We chose the free-standing MoS2/CNT–G hybrid and a commercial Pt/C catalyst (E-TEK) pasted on to a glassy carbon electrode as references to conduct parallel experiments. Fig. 5b shows the polarization curves of these three catalysts without correction for ohmic potential drop (iR) losses. The onset potential whose absolute value equals the overpotential (η) in the HER was close to zero for the Pt/C catalyst, suggesting it had high activity. The η of the MoS2/N-CNT–G catalyst was 0.08 V which was a lot lower than the MoS2/CNT–G catalyst (0.15 V). In addition, this η was also lower than some MoS2 catalysts that were supported on pristine CNTs (0.09 V)20 or graphene (0.1 V),22 suggesting that using a N-doping support could efficiently improve the activity of MoS2 catalysts for the HER. Here N-doping provided an electron-rich surface for the N-CNTs, allowing the seamlessly anchored MoS2 to create highly electronegative sites to catalyze the HER.23,24,28 But our catalyst was slightly more inactive than the MoSx/NCNT forest hybrid catalyst (η = 0.075 V) reported by Kim et al.,25 probably due to their MoSx having a higher S/Mo ratio or a more amorphous structure.
 |
| Fig. 5 (a) Photograph of the free-standing MoS2/N-CNT–G catalyst tested in a three-electrode setup; (b) polarization curves and (c) Tafel plots of the Pt/C, MoS2/CNT–G and MoS2/N-CNT–G (1 : 3) catalysts; (d) polarization curves and (e) Tafel plots of the catalysts made from different initial precursor mass ratios of 5 : 1, 1 : 3, 1 : 6 and 1 : 10. (f) Cycling test of the MoS2/N-CNT–G (1 : 3) catalyst for 1000 cycles. | |
The Tafel plots derived from the polarization curves were used to understand the catalytic process. The Tafel slope (b) in the linear regions of the Tafel plots (Fig. 5c) were calculated through fitting the Tafel equation (η = b
log
j + a, where j is the current density).38 The calculated b was 30.2, 43.5, and 48.9 mV decade−1 for the Pt/C, MoS2/N-CNT–G and MoS2/CNT–G catalysts. In general, there are three possible steps involved in the HER in acidic media, including the primary discharge step, the electrochemical desorption step and the recombination step (also named as the Volmer, Heyrovsky and Tafel reactions, respectively).39 These three steps have a distinguishable b with a value of about 120, 40 and 30 mV decade−1, respectively. As a result, we can roughly deduce which is the rate-limiting step according to the calculated b. In this case, the Pt/C catalyst gave a b of 30.2 mV decade−1, suggesting that the recombination step was the rate-limiting step of the HER. For the MoS2/N-CNT–G catalyst, the electrochemical desorption was the rate-limiting step because its b was 43.5 mV decade−1. In addition, the HER catalyzed by it belonged to the Volmer–Heyrovsky mechanism, which was consistent with other types of MoS2-based electrocatalysts.21,25
With the aim of optimizing the catalyst, three control experiments were also conducted with different N-CNT–G
:
(NH4)2MoS4 mass ratios of 5
:
1, 1
:
6 and 1
:
10. From the SEM images shown in Fig. 2b and S3,† it can be observed that the thickness of MoS2 layer decorated on the N-CNTs increased with decreasing N-CNT–G
:
(NH4)2MoS4 mass ratio. As a result, the catalyst prepared with the 5
:
1 mass ratio had the lowest MoS2 loading while the one prepared with the 1
:
10 mass ratio had the highest loading. The compositions of these catalysts were analyzed by XPS as listed in Table S1.† The mole ratio of S to Mo for each sample varied from 2.4 to 2.6, similar to that of the MoS2/N-CNT–G (1
:
3) catalyst (2.6). Their HER activities were tested in the same environment and are presented in Fig. 5d. The polarization curves indicated that these catalysts showed a lower HER activity than the MoS2/N-CNT–G (1
:
3) catalyst, which was further supported by their Tafel slopes (Fig. 5e). Since all of the catalysts had a similar S/Mo mole ratio, the factors that could have possibly affected the activity were the loading and the structure of MoS2 on the support. For example, the low loading of the catalyst prepared with the 5
:
1 mass ratio could not offer enough active sites while high loading for the catalyst (1
:
10) led to agglomeration of the MoS2 nanoparticles. Both of them were harmful for catalyzing the HER. From the structural evolution of these supported MoS2 catalysts, we could deduce that the nanosheet structure of the catalysts with the 1
:
3 mass ratio was beneficial for the exposure of active sites, and together with the proper loading amount, gave the highest activity for the HER.
Stability is also a significant property for HER catalysts. Our catalyst MoS2/N-CNT–G (1
:
3) was assessed by the cyclic voltammetry method, cycling 1000 times from −0.7 to 0.3 V at 100 mV s−1 (Fig. 5f). Compared with the initial polarization curve, the last one had almost unchanged features, indicating the strong interaction between the MoS2 nanosheets and the N-CNT–G foam. Hence the MoS2/N-CNT–G catalyst demonstrated good performance even in the free-standing configuration. After the electrochemical performance test, the used MoS2/N-CNT–G (1
:
3) catalyst was characterized by SEM (Fig. S5†). The morphology was similar to the primary material, further supporting the high stability of this catalyst.
4. Conclusions
To prepare a free-standing high performance MoS2-based electrocatalyst for the HER, CNT–graphene hybrid foams were used as a catalyst support and N-doping was applied to enhance the activity. With the assistance of N-doping, MoS2 nanosheets were uniformly and tightly anchored on N-CNTs and formed a one-dimensional structure on the graphene foams. The as-prepared MoS2/N-CNT–G catalyst showed a high activity and stability for the HER due to the additional electronegativity contributed by the N-doping. The good performance given by the free-standing configuration suggests that N-doped CNT–graphene hybrid foam is a potential catalyst support for the HER by optimizing active species.
Acknowledgements
This work is jointly supported by the Ministry of Education of China (no. IRT1148), NSFC (51402155), Priority Academic Program Development of Jiangsu Higher Education Institutions (YX03001), Jiangsu National Synergistic Innovation Center for Advanced Materials (SICAM), Synergistic Innovation Center for Organic Electronics and Information Displays, Jiangsu Provincial NSF (BK20141424).
Notes and references
- M. S. Dresselhaus and I. L. Thomas, Nature, 2001, 414, 332–337 CrossRef CAS PubMed.
- J. A. Turner, Science, 2004, 305, 972–974 CrossRef CAS PubMed.
- L. Cheng, W. J. Huang, Q. F. Gong, C. H. Liu, Z. Liu, Y. G. Li and H. J. Dai, Angew. Chem., Int. Ed., 2014, 53, 7860–7863 CrossRef CAS PubMed.
- J. Kibsgaard, Z. B. Chen, B. N. Reinecke and T. F. Jaramillo, Nat. Mater., 2012, 11, 963–969 CrossRef CAS PubMed.
- H. B. Gray, Nat. Chem., 2009, 1, 7 CrossRef CAS PubMed.
- M. G. Walter, E. L. Warren, J. R. McKone, S. W. Boettcher, Q. Mi, E. A. Santori and N. S. Lewis, Chem. Rev., 2010, 110, 6446–6473 CrossRef CAS PubMed.
- W. Cui, N. Y. Cheng, Q. Liu, C. J. Ge, A. M. Asiri and X. P. Sun, ACS Catal., 2014, 4, 2658–2661 CrossRef CAS.
- C. J. Ge, P. Jiang, W. Cui, Z. H. Pu, Z. C. Xing, A. M. Asiri, A. Y. Obaid, X. P. Sun and J. Tian, Electrochim. Acta, 2014, 134, 182–186 CrossRef CAS PubMed.
- Z. C. Xing, Q. Liu, A. M. Asiri and X. P. Sun, Adv. Mater., 2014, 26, 5702–5707 CrossRef CAS PubMed.
- W. Cui, Q. Liu, Z. C. Xing, A. M. Asiri, K. A. Alamry and X. P. Sun, Appl. Catal., 2015, 164, 144–150 CrossRef CAS PubMed.
- H. Vrubel and X. Hu, Angew. Chem., Int. Ed., 2012, 51, 12703–12706 CrossRef CAS PubMed.
- W. Cui, C. J. Ge, Z. C. Xing, A. M. Asiri and X. P. Sun, Electrochim. Acta, 2014, 137, 504–510 CrossRef CAS PubMed.
- Z. H. Pu, Q. Liu, A. M. Asirib, Y. L. Luo, X. P. Sun and Y. Q. He, Electrochim. Acta, 2015, 168, 133–138 CrossRef CAS PubMed.
- Z. H. Pu, Q. Liu, A. M. Asiri, A. Y. Obaid and X. P. Sun, J. Power Sources, 2014, 263, 181–185 CrossRef CAS PubMed.
- Y. Yang, H. L. Fei, G. D. Ruan, C. S. Xiang and J. M. Tour, Adv. Mater., 2014, 26, 8163–8168 CrossRef CAS PubMed.
- J. Kibsgaard, T. F. Jaramillo and F. Besenbacher, Nat. Mater., 2014, 6, 248–253 CAS.
- B. Hinnemann, P. G. Moses, J. Bonde, K. P. Jørgensen, J. H. Nielsen, S. Horch, I. Chorkendorff and J. K. Nørskov, J. Am. Chem. Soc., 2005, 127, 5308–5309 CrossRef CAS PubMed.
- T. F. Jaramillo, K. P. Jørgensen, J. Bonde, J. H. Nielsen, S. Horch and I. Chorkendorff, Science, 2007, 317, 100–102 CrossRef CAS PubMed.
- J. Kibsgaard, Z. B. Chen, B. N. Reinecke and T. F. Jaramillo, Nat. Mater., 2012, 11, 963–969 CrossRef CAS PubMed.
- Y. Yan, X. M. Ge, Z. H. Liu, J. Y. Wang, J. M. Lee and X. Wang, Nanoscale, 2013, 5, 7768–7771 RSC.
- H. Zhu, F. L. Lyu, M. L. Du, M. Zhang, Q. F. Wang, J. M. Yao and B.C. Guo, ACS Appl. Mater. Interfaces, 2014, 6, 22126–22137 CAS.
- Y. G. Li, H. L. Wang, L. M. Xie, Y. Y. Liang, G. S. Hong and H. J. Dai, J. Am. Chem. Soc., 2011, 133, 7296–7299 CrossRef CAS PubMed.
- K. P. Gong, F. Du, Z. H. Xia, M. Durstock and L. M. Dai, Science, 2009, 323, 760–764 CrossRef CAS PubMed.
- S. J. Jiang, Y. W. Ma, G. Q. Jian, H. S. Tao, X. Z. Wang, Z. Hu and Y. Chen, Adv. Mater., 2009, 21, 4953–4956 CrossRef CAS PubMed.
- D. J. Li, U. N. Maiti, J. Lim, D. S. Choi, W. J. Lee, Y. Oh, G. Y. Lee and S. O. Kim, Nano Lett., 2014, 14, 1228–1233 CrossRef CAS PubMed.
- G. Y. Zhu, Z. He, J. Chen, J. Zhao, X. M. Feng, Y. W. Ma, Q. L. Fan, L. H. Wang and W. Huang, Nanoscale, 2014, 6, 1079–1085 RSC.
- X. C. Dong, J. Chen, Y. W. Ma, J. Wang, M. B. Chan-Park, X. M. Liu, L. H. Wang and P. Chen, Chem. Commun., 2012, 48, 10660–10662 RSC.
- Y. W. Ma, L. Y. Sun, W. Huang, L. R. Zhang, J. Zhao, Q. L. Fan and W. Huang, J. Phys. Chem. C, 2011, 115, 24592–24597 CAS.
- L. F. Lai, J. R. Potts, D. Zhan, L. Wang, C. K. Poh, C. H. Tang, H. Gong, Z. X. Shen, J. Y. Lin and R. S. Ruoff, Energy Environ. Sci., 2012, 5, 7936–7942 CAS.
- J. Deng, W. T. Yuan, P. J. Ren, Y. Wang, D. H. Deng, Z. Zhang and X. H. Bao, RSC Adv., 2014, 4, 34733–34738 RSC.
- S. Han, Y. R. Zhao, Y. P. Tang, F. Z. Tan, Y. S. Huang, X. L. Feng and D. Q. Wu, Carbon, 2014, 81, 203–209 CrossRef PubMed.
- K. Chang and W. X. Chen, ACS Nano, 2011, 5, 4720–4728 CrossRef CAS PubMed.
- S. Zhang, X. B. Yu, H. L. Yu, Y. J. Chen, P. Gao, C. Y. Li and C. L. Zhu, ACS Appl. Mater. Interfaces, 2014, 6, 21880–21885 CAS.
- S. Qin, W. W. Lei, D. Liu and Y. Chen, Sci. Rep., 2014, 4, 7582 CrossRef CAS PubMed.
- J. F. Xie, J. J. Zhang, S. Li, F. Grote, X. D. Zhang, H. Zhang, R. X. Wang, Y. Lei, B. C. Pan and Y. Xie, J. Am. Chem. Soc., 2013, 135, 1788–17888 Search PubMed.
- K. M. Goodfellow, R. Beams, C. Chakraborty, L. Novotny and A. N. Vamivakas, Optica, 2014, 3, 149–152 Search PubMed.
- C. Y. Hu, S. Z. Zheng, C. J. Lian, F. Chen, T. W. Lu, Q. H. Hu, S. W. Duo, R. B. Zhang and C. Y. Guan, J. Mol. Catal. A: Chem., 2015, 396, 128–135 CrossRef CAS PubMed.
- B. E. Conway and B. V. Tilak, Electrochim. Acta, 2002, 47, 3571–3594 CrossRef CAS.
- W. C. Sheng, H. A. Gasteiger and Y. Shao-Horn, J. Electrochem. Soc., 2010, 157, B1529–B1536 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra05220c |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.